16
Visual working memory for image statistics Jonathan D. Victor * , Mary M. Conte Department of Neurology and Neuroscience, Weill Medical College of Cornell University, 1300 York Avenue, New York, NY 10021, USA Received 20 May 2003; received in revised form 3 November 2003 Abstract To define the role of statistical features of images in visual working memory, we compared the ability of subjects ðN ¼ 6Þ to identify changes in arrays of black and white checks when these changes altered some aspect of their statistical structure, versus when these changes did not. Alteration of luminance statistics or local higher-order statistics improved performance, but alteration of the degree of bilateral symmetry did not. The dependence of performance on the degree of statistical change indicated that statistical information was represented in a graded, rather than categorical, fashion. Ó 2003 Elsevier Ltd. All rights reserved. Keywords: Symmetry; Isodipole 1. Introduction Early vision segments the retinal image into objects and represents these objects in a manner in which they can be recognized. Individual features such as lines and edges play a role in these processes, but often the sta- tistics of visual images are at least as important. For example, in complex images, including naturalistic ones, only a small number of contrast contours represent object boundaries, and consequently objects are more reliably defined by discontinuities in image statistics, rather than by isolated features (Julesz, 1981a; Marr, 1982). Despite the impressive ability of the visual system to make use of scene statistics, much previous work (Caelli & Julesz, 1978; Caelli, Julesz, & Gilbert, 1978; Julesz, Gilbert, Shepp, & Frisch, 1973; Julesz, Gilbert, & Victor, 1978; Victor & Conte, 1991, 1996), which has focused on texture discrimination and segmentation, indicates that this statistical processing is limited and specific. In natural viewing, segmentation of an image proceeds along with an analysis of surface composition. This analysis (e.g., sand versus wood versus stone) is doubtless based on image statistics (Cho, Yang, & Hallett, 2000), rather than a pixel-by-pixel match with an exemplar, and thus, represents another situation in which image statistics play a crucial role. To determine the extent to which visual image sta- tistics are available for such processes as distinct from spatial segmentation or explicit discrimination tasks, we made use of a visual working memory task introduced by Cornelissen and Greenlee (2000). These authors showed that human subjects are able to determine whether a random array has changed over a brief time interval, but performance on this task is poor. Most likely, the poor performance on this task reflects the limited capacity of visual working memory for encoding and/or representing such images on a pixel-by-pixel basis. Consequently, if alteration in an array could be detected by a change in image statistics, then perfor- mance might improve dramatically. That is, we can as- say the extent to which visual working memory makes use of image statistics by determining the extent to which introduction of specific kinds of statistical struc- ture affect performance on this task. Since there are far too many image statistics to at- tempt a rigorously comprehensive analysis (Cho et al., 2000; Harvey & Gervais, 1981), we adopt a ‘‘survey’’ strategy motivated by physiological principles and pre- vious work with texture discrimination and segmenta- tion. We will therefore consider exemplars of three classes of visual statistics (see Fig. 2 for examples). The first class consists of image statistics that influence the mean activity of a population of retinal ganglion cells. This includes luminance and second-order corre- lation structure; we will use luminance (the fraction of white checks) as an exemplar of this class (Fig. 2(A)). A * Corresponding author. Tel.: +1-212-746-2343; fax: +1-212-746- 8984. E-mail address: [email protected] (J.D. Victor). 0042-6989/$ - see front matter Ó 2003 Elsevier Ltd. All rights reserved. doi:10.1016/j.visres.2003.11.001 Vision Research 44 (2004) 541–556 www.elsevier.com/locate/visres

Visual working memory for image statistics

Embed Size (px)

Citation preview

Page 1: Visual working memory for image statistics

Vision Research 44 (2004) 541–556

www.elsevier.com/locate/visres

Visual working memory for image statistics

Jonathan D. Victor *, Mary M. Conte

Department of Neurology and Neuroscience, Weill Medical College of Cornell University, 1300 York Avenue, New York, NY 10021, USA

Received 20 May 2003; received in revised form 3 November 2003

Abstract

To define the role of statistical features of images in visual working memory, we compared the ability of subjects ðN ¼ 6Þ to

identify changes in arrays of black and white checks when these changes altered some aspect of their statistical structure, versus

when these changes did not. Alteration of luminance statistics or local higher-order statistics improved performance, but alteration

of the degree of bilateral symmetry did not. The dependence of performance on the degree of statistical change indicated that

statistical information was represented in a graded, rather than categorical, fashion.

� 2003 Elsevier Ltd. All rights reserved.

Keywords: Symmetry; Isodipole

1. Introduction

Early vision segments the retinal image into objects

and represents these objects in a manner in which they

can be recognized. Individual features such as lines and

edges play a role in these processes, but often the sta-

tistics of visual images are at least as important. For

example, in complex images, including naturalistic ones,

only a small number of contrast contours representobject boundaries, and consequently objects are more

reliably defined by discontinuities in image statistics,

rather than by isolated features (Julesz, 1981a; Marr,

1982). Despite the impressive ability of the visual system

to make use of scene statistics, much previous work

(Caelli & Julesz, 1978; Caelli, Julesz, & Gilbert, 1978;

Julesz, Gilbert, Shepp, & Frisch, 1973; Julesz, Gilbert, &

Victor, 1978; Victor & Conte, 1991, 1996), which hasfocused on texture discrimination and segmentation,

indicates that this statistical processing is limited and

specific. In natural viewing, segmentation of an image

proceeds along with an analysis of surface composition.

This analysis (e.g., sand versus wood versus stone) is

doubtless based on image statistics (Cho, Yang, &

Hallett, 2000), rather than a pixel-by-pixel match with

an exemplar, and thus, represents another situation inwhich image statistics play a crucial role.

* Corresponding author. Tel.: +1-212-746-2343; fax: +1-212-746-

8984.

E-mail address: [email protected] (J.D. Victor).

0042-6989/$ - see front matter � 2003 Elsevier Ltd. All rights reserved.

doi:10.1016/j.visres.2003.11.001

To determine the extent to which visual image sta-tistics are available for such processes as distinct from

spatial segmentation or explicit discrimination tasks, we

made use of a visual working memory task introduced

by Cornelissen and Greenlee (2000). These authors

showed that human subjects are able to determine

whether a random array has changed over a brief time

interval, but performance on this task is poor. Most

likely, the poor performance on this task reflects thelimited capacity of visual working memory for encoding

and/or representing such images on a pixel-by-pixel

basis. Consequently, if alteration in an array could be

detected by a change in image statistics, then perfor-

mance might improve dramatically. That is, we can as-

say the extent to which visual working memory makes

use of image statistics by determining the extent to

which introduction of specific kinds of statistical struc-ture affect performance on this task.

Since there are far too many image statistics to at-

tempt a rigorously comprehensive analysis (Cho et al.,

2000; Harvey & Gervais, 1981), we adopt a ‘‘survey’’

strategy motivated by physiological principles and pre-

vious work with texture discrimination and segmenta-

tion. We will therefore consider exemplars of three

classes of visual statistics (see Fig. 2 for examples). Thefirst class consists of image statistics that influence

the mean activity of a population of retinal ganglion

cells. This includes luminance and second-order corre-

lation structure; we will use luminance (the fraction of

white checks) as an exemplar of this class (Fig. 2(A)). A

Page 2: Visual working memory for image statistics

542 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

second class of visual statistics requires cortical analysis

for extraction, but it suffices that this analysis occur

within a local region. We will use fourth-order correla-

tions, as manifest by the ‘‘even and odd’’ isodipole

textures (Julesz et al., 1978) as an exemplar of this class

(Fig. 2(B)). Sensitivity of cortical (Purpura, Victor, &

Katz, 1994) but not lateral geniculate neurons (Victor,

1986) to these statistics has been demonstrated experi-mentally. A third class of statistics can only be extracted

via cortical analysis that is long-range. We will use

bilateral symmetry, widely considered a salient and

important visual feature (Attneave, 1954; Tyler, 1995;

Wenderoth, 1994), as an exemplar of this class (Fig.

2(C)–(E)). Textures that isolate the first two kinds of

statistical structure can readily be constructed by

homogeneous Markov random fields (Zhu, Wu, &Mumford, 1998), while the third cannot. Nevertheless,

as described below, each of these kinds of structure can

be introduced into otherwise random arrays in a quan-

titative and graded fashion, allowing us to measure their

effects independently and on a common footing. More-

over, because each kind of structure can be precisely

quantified, an information-theoretic analysis can be

used to compare the intrinsic difficulty of the psycho-physical tasks. As shown below, only the first two

classes of image statistics appear to be used in visual

working memory, even though bilateral symmetry is

widely considered to be visually salient.

Fig. 1. A diagram of a typical trial. The subject’s task is to determine

which of the four arrays in S1 has changed in S2.

2. A categorical representation?

If indeed image statistics are used to identify or

classify surface materials (Cho et al., 2000), one might

speculate that they play a role in object identification

analogous to that of color. A distinctive feature of color

perception, both in discrimination (Bornstein & Korda,1984; Wandell, 1985) and visual memory tasks (Amano,

Uchikawa, & Kuriki, 2002) is that the physical contin-

uum of color space is represented in a categorical

manner. In a categorical representation, performance

depends primarily on where these stimuli are located

with respect to one or more boundaries in a perceptual

space (Berlin & Kay, 1969; Bornstein & Korda, 1984;

Wandell, 1985). Stimuli are distinctive on opposite sidesof such a boundary. Accuracy on a discrimination task

is enhanced and reaction time decreases on trials in

which stimuli are drawn from opposite sides of a cate-

gory boundary, compared to trials in which both stimuli

are drawn from the same category. The alternative is a

graded representation, in which the parametric differ-

ence between two stimuli, not their position relative to a

perceptual boundary, is the main determinant of per-formance. For those classes of image statistics that are

used in visual working memory, our approach allows us

to ask whether this representation is categorical or

graded. Although the analogy to color and the role of

textures in material or surface classification might sug-

gest a categorical representation, the picture that emer-

ges is largely a graded one.

3. Methods

3.1. Subjects

Studies were conducted in six normal subjects (two

male, four female), ages 30–54. Other than author MC,

the remaining subjects were na€ııve to the purpose of the

experiments. Subjects were practiced psychophysicalobservers in a related task involving targets in the same

positions relative to fixation (Victor & Conte, 2001), and

had visual acuities (corrected if necessary) of 20/20 or

better.

3.2. Stimuli

The stimulus frame S1 (Fig. 1) consists of four arrays

of checks on a mean gray background. The arrays werepositioned along the cardinal axes, with centers 200 min

from fixation. In most experiments, each array sub-

tended 160 min and contained 64 (8 · 8) contiguous

checks, each of which was either black or white and

subtended 20 min. The stimulus frame S2 also consisted

of four arrays, three of which were identical to those in

the S1 frame of the trial. The target array, determined at

random, differed from the corresponding array in S1 bya contrast inversion of 16 of the 64 checks. In other

experiments, the arrays were of equal size but contained

only 16 (4 · 4) checks, each subtending 40 min. In these

experiments, the target array differed from the corre-

sponding array in S1 by inversion of eight of the 16

checks. The number of checks that differed between S1

and S2 was chosen so that performance was neither at

chance nor ceiling, and at approximately the same levelfor the two array sizes.

Page 3: Visual working memory for image statistics

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 543

The main experimental manipulation consisted of the

assignment of luminance values to the checks. For each

experiment, a particular kind of statistical structure was

introduced into the arrays: luminance bias, higher-order

statistical structure (the ‘‘isodipole’’ textures), or sym-

metry. In each case, the strength of the statistical

structure was parameterized by a quantity c, where

c ¼ 0 denotes a maximally random assignment, andc ¼ 1 (or c ¼ �1) denotes a maximally structured

assignment. In the experiments that examined lumi-

nance statistics, an array corresponding to a value c had1þc2

of its checks white, and 1�c2

of its checks black. In the

experiments that examined higher-order statistics, c ¼ 1

corresponded to a maximally ‘‘even’’ texture, while

c ¼ �1 corresponded to a maximally odd texture. In the

symmetry experiments, c ¼ 1 corresponded to a texturein which all pairs of checks that were related by the

symmetry axis were matched in luminance, while c ¼ �1

corresponded to a texture in which all such pairs were

opposite in luminance. Appendix A provides details the

construction of these stimuli, including the precise defi-

nition of the textures for intermediate values of c.Within an experiment, trials were constructed with

one or more pairs of c-values, as follows. To construct atrial based on the pair ðclow; chighÞ, each of the four ar-

rays in S1 was constructed either with c ¼ clow or with

c ¼ chigh. The c-value of the target array in S2 was also

randomly assigned to one of these two values of c. Thus,for each pair of c-values, there were 128¼ 25 · 4 varietiesof stimuli, since for each of the five arrays (four in S1,

and the target in S2) there were two possible values for

c, and the target could occur in any of four locations.Each of these varieties was presented the same number

of times. In exactly half of the trials, designated ‘‘dif-

ferent statistics’’ trials below, the c-value of the target

changed from S1 to S2 by an amount Dc ¼ jchigh � clowj;in half of these trials, c changed by þDc (increasing fromclow to chigh), and in half it changed by �Dc (decreasing

from chigh to clow). In the other half of the trials, desig-

nated ‘‘same statistics’’ trials, the c-value of the targetremained at either clow or chigh. Thus, neither the c-valueschosen for the arrays in S1, nor the c-values chosen in

S2, provided a cue as to which array was the target. As

described in the Appendix A, it is possible to construct

sets of arrays that not only have the requisite values of c,but also differ in a specific number of checks, as required

by the experimental design.

3.3. Apparatus

The above visual stimuli were produced on a Sony

Multiscan 17seII (1700diagonal) monitor, with signals

driven by a PC-controlled Cambridge Research VSG2/3graphics processor programmed in Delphi II to display

precomputed maps (generated in Matlab) for specified

periods of time. The resulting 768 · 1024 pixel display

had a mean luminance of 47 cd/m2, a refresh rate of 100

Hz and subtended 11 · 15 deg (approximately 1 min/

pixel) at the viewing distance of 114 cm. The intensity

versus voltage behavior of the monitor was linearized by

photometry and lookup table adjustments provided by

VSG software. Stimulus contrast was 1.0.

3.4. Procedure

Essentially, our experimental design is a modification

of the Cornelissen and Greenlee (2000) visual working

memory task, in which (a) four stimuli are presented

simultaneously, and (b) statistical cues are intentionallyintroduced. All experiments were organized as a se-

quence of 4-alternative forced choice trials, whose

common features are as follows (Fig. 1). After binocular

fixation on a uniform gray background, the subject

initiated a trial via a button-press on a Cambridge Re-

search CT3 response box 300 ms later, a stimulus (S1,

described in detail above) appeared, consisting of four

arrays of checks, surrounding a central ‘‘X’’ subtendingapproximately 30 min. After presentation of S1 for 600

ms, the display returned to mean luminance for 200 ms,

following which a second stimulus S2 (described above)

appeared, containing a ‘‘target’’ that differed from the

corresponding array in S1. After presentation of S2 for

200 ms, a mask was presented for 500 ms, consisting of a

full-field random checkerboard whose checks were half

as large (linear dimension) as those in S1 and S2. Thesubject’s task was to identify the target array via a

button-press on a response box with four buttons,

positioned corresponding to the stimulus arrays. Sub-

jects were instructed to maintain central fixation and to

respond as quickly as possible, but not to compromise

accuracy for speed. Responses and reaction times

(measured with respect to the onset of S2) were collected

via the Delphi II display software. Trials in which thesubject responded before the onset of S2, or after 8000

ms, were discarded and repeated.

An experimental session consisted of a single block of

either 512 or 640 trials (4 or 5 · the 128 varieties of trials

that were required to examine each kind of statistical

structure and each pair of c-values, presented in random

order). In Experiment I, one pair of c-values was

examined for each kind of statistical structure. InExperiment II, four or five pairs of c-values were

examined for each kind of statistical structure. To

accumulate a sufficient number of trials for each c-valuepair in Experiment II, four sessions (on separate days)

were required. In Experiment I, subjects were shown

paper exemplars of trials with the maximal level of

statistical structure at the beginning of each session. In

Experiment II, subjects were shown exemplars thatspanned the c-values to be used. In both cases, subjects

were informed that the trials would be similar to these

exemplars. For data analysis in Experiment II, results

Page 4: Visual working memory for image statistics

544 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

from each subject were pooled across sessions. Thus, the

critical comparisons between c-values in Experiment II

were based on trials run in parallel.

Fig. 2. The five kinds of statistical manipulations used in Experiment I

(left), and a summary of performance (right). Fraction correct is

pooled across subjects, and the error bars represent 95% confidence

limits of the pooled values, based on binomial statistics. The size of the

array and the number of checks that change is indicated by symbol

shape: squares, eight checks change within an array of 16 checks; tri-

angles, 16 checks change within an array of 64 checks. Open symbols:

‘‘different statistics’’ trials; filled symbols: ‘‘same statistics’’ trials. (A)

Luminance statistics, (B) isodipole statistics, (C) vertical symmetry

versus absence of symmetry, (D) vertical symmetry versus horizontal

symmetry, (E) vertical symmetry versus contrast-inverted vertical

symmetry.

4. Results

4.1. Experiments I: What kinds of image statistics are

available in visual working memory?

In these five experiments, we examined the effects of

large variations of statistics on visual working memory.

We examined two kinds of local statistics (luminanceand isodipole) and three examples of the third class of

statistics (global symmetry). In each experiment, the

target array either was constructed with statistics that

differed by a large amount Dc ¼ jchigh � clowj along one

of these axes, or not at all.

Fig. 2 illustrates the stimuli and summarizes the re-

sults from the five sub-experiments. For the luminance

experiment (Fig. 2(A)), we used chigh ¼ 0:25 andclow ¼ �0:25, so Dc ¼ 0:5 in the ‘‘different statistics’’

trials. This led to a substantially higher fraction correct,

compared to the fraction correct in the ‘‘same statistics’’

trials, in each of the six subjects. On average, fraction

correct improved from 0.69 to 0.89 for the 16-check

arrays in which eight checks changed, and from 0.56 to

0.84 for the 64-check arrays in which 16 checks changed.

The improvement in performance was highly statisti-cally significant (p < 10�4 for each subject and each ar-

ray size individually, two-tailed v2). Here and below, we

only consider differences significant if there are consis-

tent differences in most subjects, when data are analyzed

individually (without Bonferroni correction).

For the isodipole experiment (Fig. 2(B)), we used

chigh ¼ 1 (maximally ‘‘even’’) and clow ¼ �1 (‘‘odd’’), so

Dc ¼ 2 in the ‘‘different statistics’’ trials. As with lumi-nance statistics, there was a substantial increase in

fraction correct in the ‘‘different statistics’’ trials. On

average, fraction correct improved from 0.79 to 0.86 for

the 16-check arrays in which eight checks changed, and

from 0.58 to 0.83 for the 64-check arrays in which 16

checks changed. The improvement in performance was

highly statistically significant (p < 0:02 for four of six

subjects for the 16-check array, and p < 10�4 for eachsubject for the 64-check array). The finding that there is

a smaller improvement for the 16-check array than for

the 64-check array is entirely due to the fact that per-

formance in the ‘‘same statistics’’ trials for 16-check

arrays was higher than for 64-check arrays.

Increases in fraction correct were accompanied by

decreases in reaction time (calculated by averaging

across trials for which the response was correct). Inluminance experiments, the reaction time decreases were

large and statistically significant for both array sizes (16-

check array: mean RT decrease across subjects, 144 ms;

p < 0:05 for five of six subjects by one-tailed t-test; 64-check array: mean RT decrease across subjects, 191 ms;

p < 0:01 for all six subjects). For the isodipole experi-

ments, the change in reaction time was minimal for the

16-check array, but large for the 64-check array (16-

check array: mean RT decrease across subjects, 20 ms;

p < 0:05 for one subject; 64-check array: mean RT de-

crease across subjects, 80 ms; p < 0:01 for four of six

subjects). This closely paralleled the pattern of findingsfor fraction correct. In sum, these experiments show that

change in an array is easier to detect when its luminance

statistics (Fig. 2(A)) or isodipole statistics (Fig. 2(B))

change.

For the symmetry experiments, the results were quite

different. Fig. 2(C) shows an experiment in which some

arrays had perfect bilateral symmetry along the vertical

axis ðchigh ¼ 1Þ, and others were random ðclow ¼ 0Þ.There was no suggestion of a significant difference in

Page 5: Visual working memory for image statistics

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 545

performance between the ‘‘different statistics’’ trials

ðDc ¼ 1Þ and the ‘‘same statistics’’ trials (p > 0:05 in

each subject for each array size; p ¼ 0:14 pooled across

subjects for the 16-check arrays; p ¼ 0:49 pooled across

subjects for the 64-check arrays). Reaction times were

minimally decreased in the ‘‘different statistics’’ trials

(16-check array: mean RT decrease across subjects, 40

ms; p < 0:05 for two of six subjects; 64-check array:mean RT decrease across subjects, 18 ms; p < 0:01 for

one subject).

We sought to increase the magnitude of this effect in

two ways. In the experiment of Fig. 2(D), all arrays were

fully symmetric, but some had a vertical symmetry axis,

while others had a horizontal symmetry axis. Overall

performance was somewhat better, but again there was

no difference in performance between the ‘‘differentstatistics’’ trials and the ‘‘same statistics’’ trials (p > 0:05in each subject for each array size; p ¼ 0:34 pooled

across subjects for the 16-check arrays; p ¼ 0:13 pooled

across subjects for the 64-check arrays). In the experi-

ment of Fig. 2(E), we implemented the maximum pos-

sible modulation of vertical symmetry, by using arrays

that either had perfect perfect bilateral symmetry along

the vertical axis ðchigh ¼ 1Þ, or those in which checksrelated by the mirror pairing mismatched in luminance

ðclow ¼ �1Þ. As described in Appendix A, such stimuli

could only be realized with the 16-check array size.

While there was a statistically significant difference in

performance between the ‘‘different statistics’’ trials and

the ‘‘same statistics’’ trials (p < 10�4 across all subjects),

this effect was not robust (significant at p < 0:05 in only

two subjects, MC and EM). Moreover, this small dif-ference might be due to the fact that all of the stimuli

with mismatch along the mirror axis ðclow ¼ �1Þ neces-sarily had a vertical contrast contour at the midline,

while the stimuli with perfect vertical symmetry

ðchigh ¼ 1Þ necessarily had no contour at this location.

Correspondingly, there was no consistent effect of a

symmetry change on reaction time (vertical versus hor-

izontal symmetry: RT decrease of 22 ms for the 16-checkarray and RT increase of 15 ms for the 64-check array;

symmetry versus mismatch: RT decrease of 11 ms; only

three of 18 comparisons significant within subjects at

p < 0:05). Thus, in contrast to our findings with local

statistics (Fig. 2(A) and (B)), gain, loss, or change of

bilateral symmetry did not appear to provide a useful

cue in this visual working memory task.

4.1.1. Role of spatial differences in image statistics

Differences between the statistical structure of the

arrays might influence performance not only via aiding

representation in working memory, but also via spatial

differences within S1 or S2––for example, by guidingattention. The statistical structure of each of the four

arrays (c ¼ chigh versus c ¼ clow) was assigned indepen-

dently, and with equal probability. Thus, on one quarter

of the trials, a triplet of the S1 arrays had c ¼ chigh and

the remaining array (a ‘‘singleton’’) had c ¼ clow, and in

a separate quarter of the trials, a triplet of S1 arrays had

c ¼ clow and the remaining singleton array had c ¼ chigh.If the singleton array drew spatial attention by virtue of

their distinctive statistics (i.e., elicited ‘‘pop-out’’), one

might expect that performance would be enhanced on

trials in which the target was also a singleton array in S1or S2 (Joseph & Optican, 1996; Treisman, 1982).

To identify the role played by spatial differences in

image statistics, we separated the trials into three

groups: those in which the target was a singleton in S1

(here designated ‘‘pop-out in S1’’), those in which the

target was a singleton in S2 (here designated ‘‘pop-out in

S2’’), and the remaining trails, in which the target was

not a singleton in either S1 or S2 (here designated ‘‘nopop-out’’). We use these terms as a shorthand, and not

to imply a mechanism. They indicate not only the

presence or absence of a singleton array, but specifically

whether the singleton is also the target.

We found that for some of the statistical classes,

fraction correct was higher in the pop-out trials than in

the ‘‘no pop-out’’ trials. For the luminance experiments

based on 64-check arrays, average fraction correct was0.73 in the ‘‘pop-out in S1’’ trials, 0.76 in the ‘‘pop-out

in S2’’ trials, and 0.60 in the ‘‘no pop-out’’ trials. For

isodipole experiments, the corresponding fractions cor-

rect were 0.75 in the ‘‘pop-out in S1’’ trials, 0.78 in the

‘‘pop-out in S2’’ trials, and 0.61 in the ‘‘no pop-out’’

trials. Differences between the pop-out and ‘‘no pop-

out’’ trials were highly significant (p < 10�4 pooled

across subjects for luminance and for isodipole statis-tics); differences between ‘‘pop-out in S1’’ and ‘‘pop-out

in S2’’ were not significant (p > 0:2 pooled across sub-

jects for luminance and isodipole classes).

Reaction time differences among these three kinds of

trials were generally small, and not statistically signifi-

cantly within subjects. Note that the number of correct

pop-out trials was typically 50–60 for each subject (out

of 64–80 pop-out trials), so small differences in reactiontimes might not reach statistical significance. Pooling

data across subjects only revealed a statistically signifi-

cant difference in RT in the luminance trials, and only

when pop-out occurred in S2. Compared with the ‘‘no

pop-out’’ trials, RT in ‘‘pop-out in S2’’ trials was re-

duced by 43 ms (16-check arrays, p < 0:02) and 49 ms

(64-check arrays, p < 0:02).In sum, an effect of spatial differences in image sta-

tistics on fraction correct was seen for both statistical

classes that showed an effect of ‘‘different statistics’’

versus ‘‘same statistics’’, and on RT for the luminance

class. However, the effect of statistical change (Fig. 2)

was larger than, and not accounted for by, this phe-

nomenon. This is seen in Fig. 3. As seen in Fig. 3(A) for

the luminance experiments, the difference in fraction

correct between the ‘‘same statistics’’ trials and in the

Page 6: Visual working memory for image statistics

0.4

1.0

no pop-out pop-out in S1 pop-out in S2

0.6

0.8

luminance

isodipole

0.4

1.0

frac

tion

corr

ect

0.6

0.8

(A)

(B)

no pop-out pop-out in S1 pop-out in S2

Fig. 3. Analysis of performance for trials in which the target’s statis-

tics can guide spatial attention by virtue of its statistics in S1, or in S2,

or in neither trial. Open symbols: ‘‘different statistics’’ trials ðDc 6¼ 0Þ;filled symbols: ‘‘same statistics’’ trials ðDc ¼ 0Þ. (A) Luminance sta-

tistics (Dc ¼ 0:5 or 0), (B) isodipole statistics (Dc ¼ 2 or 0). See text for

details on how trials were classified as ‘‘pop-out’’ and ‘‘no pop-out’’

trials. Data for 64-element arrays. Error bars as in Fig. 2.

Fig. 4. Examples of stimuli for the luminance version of Experiment

II. Each column contains three representative examples of arrays

constructed with the indicated value of the structure parameter c.Adjacent columns are separated by Dc ¼ 0:25, the amount of statistical

change used in the experiment.

546 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

‘‘different statistics’’ trials persisted when ‘‘pop-out in

S1’’, ‘‘pop-out in S2’’, and ‘‘no pop-out’’ trials were

separately analyzed. A similar pattern was seen for the

experiment based on isodipole statistics, as seen in Fig.

3(B), and in the trials based on the 16-check arrays (not

shown).

Fig. 5. Results of the luminance experiment with Dc ¼ 0:25. Open

circles: ‘‘different statistics’’ trials; filled circles: ‘‘same statistics’’ trials.

(A) Fraction correct, pooled across four subjects. The dashed arrows

indicate the ðclow; chighÞ pair associated with an example data point; the

other data points correspond to an equally separated pair of values.

(B) Reaction time, pooled across four subjects. Error bars for fraction

correct as in Fig. 2; error bars for reaction times are the means of the

95% confidence limits within each subject.

4.2. Experiments II: Structure of the perceptual space

The previous experiments established that certain

image statistics, and not just the pixel-by-pixel details of

an image, were represented in visual working memory.

Since all instances of ‘‘different statistics’’ consisted of

large changes (Dc ¼ 0:5 for luminance statistics, Dc ¼ 2

for isodipole statistics), this result leaves open thequestion of whether this representation is categorical or

graded. The next set of experiments address this issue by

examining the influence of small changes in statistics.

We restricted consideration to the 64-element arrays,

since this size provided a larger effect in Experiment I

than the 16-element arrays (Fig. 2(A) and (B)). Four of

the six subjects who participated in Experiment I pro-

vided data for all of the experiments described here. Afifth subject from Experiment I (KS) also provided data

for the first isodipole statistics experiment described

below.

4.2.1. Luminance statistics

Luminance statistics were investigated with

Dc ¼ 0:25, and stimulus pairings ranging from

ðclow; chighÞ ¼ð�0:625;�0:375Þ to ðclow; chighÞ ¼ ðþ0:375;þ0:625Þ. Examples of the arrays used are shown in Fig.

4. As seen in Fig. 5(A), fraction correct in the ‘‘different

Page 7: Visual working memory for image statistics

Fig. 7. Results of the isodipole experiment with Dc ¼ 0:4. (A) Fraction

correct, pooled across subjects. (B) Reaction time, pooled across

subjects. Plotting conventions as in Fig. 5.

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 547

statistics’’ trials was higher than that in the ‘‘same sta-

tistics’’ trials. This difference, about 0.1, was indepen-

dent of the position of the stimuli along the range of

luminance statistics. It was highly statistically significant

ðp < 10�4Þ for each pairing of c-values in data pooled

across subjects. The same pattern was seen in all indi-

vidual subjects, though not all comparisons (3, 4 or 5

out of 5) reached statistical significance ðp < 0:05Þ indata from individual subjects. Correspondingly, reac-

tion time (Fig. 5(B)) was shorter in the ‘‘different sta-

tistics’’ trials than in the ‘‘same statistics’’ trials. The

reaction time reduction, on average 54 ms, was also

independent of the position of the stimuli along the

range of luminance statistics. Reaction time changes

were of only modest statistical significance across sub-

jects (p ¼ 0:04–0.07 via one-tailed paired t-test at four ofthe five pairings, p > 0:2 at the middle pairing) and

within subjects (p < 0:05 at 2–5 of the pairings in indi-

vidual subjects), most likely because of the intersubject

variability of reaction times.

4.2.2. Isodipole statistics

A corresponding analysis for isodipole statistics, with

Dc ¼ 0:4 (Fig. 6) is shown Fig. 7. The only clear differ-

ence in fraction correct between ‘‘different statistics’’

and ‘‘same statistics’’ trials was for ðclow; chighÞ ¼ ðþ0:6;þ1Þ (p < 10�3 pooled across subjects, p < 0:05 in four of

five individual subjects). For seven of the other eight

pairs ðclow; chighÞ tested, there was a tendency, though notstatistically significant, in the same direction. Reaction

time data (Fig. 7(B)) showed no clear difference

ðp > 0:05Þ between conditions at all nine pairings, both

within and across subjects.

These data suggest that there is something unique

about the fully ‘‘even’’ ðc ¼ 1Þ stimulus, but leave open

the possibility that a small difference (that failed to reach

statistical significance) was also present at lower valuesof c. For this reason, we also measured performance for

Fig. 6. Examples of stimuli for the isodipole version of Experiment II.

Each column contains three representative examples of arrays con-

structed with the indicated value of the structure parameter c. Adjacent

columns are separated by Dc ¼ 0:2. Stimulus pairings (indicated by

arrows) were constructed from examples drawn from next-nearest

neighbors ðDc ¼ 0:4Þ.

Fig. 8. Results of the isodipole experiment with Dc ¼ 1. (A) Fraction

correct, pooled across four subjects. (B) Reaction time, pooled across

four subjects. Plotting conventions as in Fig. 5.

stimuli constructed with Dc ¼ 1. As seen in Fig. 8(A), a

change in image statistics does improve performanceacross the entire range of c (p < 0:05 at each pairing,

pooled across subjects). However, this improvement is

larger near the extreme ‘‘even’’ ðc ¼ 1Þ end, both in

terms of magnitude and statistical significance. In par-

ticular, there was an improvement in performance for

one or both of the pairings at the ‘‘even’’ end ðc ¼ 1Þ ofthe range at p < 0:05 in all four subjects, but only one

Page 8: Visual working memory for image statistics

548 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

subject’s individual data showed a comparable differ-

ence at the ‘‘odd’’ end ðc ¼ �1Þ of the range. There wasa 31 ms reduction of reaction time in the ‘‘different

statistics’’ trials across the entire range (Fig. 8(B)). In

keeping with the previously observed close parallel of

reaction time and fraction correct, the reduction in

reaction time was greater at the ‘‘even’’ end of the range

(77 ms, p < 0:05) than at the ‘‘odd’’ end (14 ms, p >0:15).

4.2.3. Dependence of performance on the statistics of the

target

We have shown that a change in the statistics of the

target between S1 and S2 leads to improved perfor-

mance, across the entire range of luminance and isodi-pole statistics. Here we ask whether there is an influence

A. luminance

fract

ion

corre

ct

target c-leve

target c

target c

0.6

0.8

1.0

-0.75 -0.5 -0.25 0 0.25 0.5 0.75

Chigh to Chigh

Clow to Clow

-0.75 -0.5 -0.25 0 0.25 0.5 0.75

Chigh to Clow

Clow to Chigh

-0.75 -0.5 -0.25 0 0.25 0.5 0.75

Chigh to Clow

Clow to Chigh

0.6

0.8

1.0

0.6

0.8

1.0

Fig. 9. Dependence of fraction correct on the c-value of the target. (A) Lumin

of c represent, respectively, the dark and light ends of the range of luminance

positive values of c represent, respectively, the odd and even ends of the ra

formance in the ‘‘same statistics’’ trials, as a function of the c-value of the tarand S2. Filled symbols: target c ¼ chigh in S1 and S2. Middle plot: performan

target in S1. Open symbols: target c ¼ clow in S1 and c ¼ chigh in S2. Filled sym

in the ‘‘different statistics’’ trials, as a function of the c-value of the target in S

positions of the points along the abscissa.

of the position of the target along the statistical range in

S1 or in S2, in addition to the effect of a change in po-

sition between S1 and S2. To make this distinction, we

separately consider the ‘‘same statistics’’ and ‘‘different

statistics’’ trials in the above experiments.

We consider the luminance experiments (Fig. 9(A))

first. For the ‘‘same statistics’’ trials (upper panel), there

is no overall trend of the fraction correct as a function ofthe c-value of the target (regression slope 0.036, p > 0:1,F -test). However, for trials in which the stimuli were at

the dark ðc < 0Þ end of the range, fraction correct was

greater ðp < 0:03Þ when the target was the brighter of

the two alternatives ðc ¼ chighÞ. Conversely, for trials inwhich the stimuli were at the bright ðc > 0Þ end of the

range, fraction correct was greater ðp < 10�4Þ when the

target was the darker of the two alternatives ðc ¼ clowÞ.

B. isodipole

l in S1 and S2

-level in S1

-level in S2

-1 -0.5 0 0.5 1

Chigh to Chigh

Clow to Clow

Chigh to Clow

Clow to Chigh

Chigh to Clow

Clow to Chigh

0.6

0.8

1.0

-1 -0.5 0 0.5 1

-1 -0.5 0 0.5 1

0.6

0.8

1.0

0.6

0.8

1.0

ance experiment of Fig. 5 with Dc ¼ 0:25. Negative and positive values

statistics. (B) Isodipole experiment of Fig. 8 with Dc ¼ 1. Negative and

nge of luminance statistics. Error bars as in Fig. 2. Upper plot: per-

get, which is identical in S1 and S2. Open symbols: target c ¼ clow in S1

ce in the ‘‘different statistics’’ trials, as a function of the c-value of thebols: target c ¼ chigh in S1 and c ¼ clow in S2. Lower plot: performance

2; same convention for symbols. The two lower plots differ only in the

Page 9: Visual working memory for image statistics

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 549

This suggests a modest contribution of a contextual

pop-out (i.e., statistics that deviate from the end of the

range typified by the pair ðclow; chighÞ that define the

block of trials).

When the statistics of the target do change, then the

target c must necessarily equal one of the two paired

values clow or chigh in S1, and the other value in S2. Thus,

the statistical change necessarily nulls any effect ofcontextual pop-out. Analysis of fraction correct as a

function of the c-level of the target in S1 (Fig. 9(A),

middle panel) or in S2 (Fig. 9(A), lower panel) show that

there is no overall dependence of performance on the

position of the target along the range (regression slopes

of 0.052 and )0.003, respectively, both p > 0:1, F -test).However, for target c-values of )0.125, 0.125, and 0.375,

fraction correct is higher for trials in which the target isdarker in S2 than in S1, than for trials in which the

target is brighter in S2 than S1 (p < 0:05 if comparison

is based on target statistics in S1, p < 0:01 if comparison

is based on target statistics in S2).

The isodipole experiments with Dc ¼ 1 (Fig. 9(B))

show a very different pattern. For the ‘‘same statistics’’

trials (upper panel), there is a small but significant in-

crease in fraction correct as a function of the c-value ofthe target (regression slope 0.049, p < 0:05, F -test).There is only one c-value, namely c ¼ 0, that is tested

both as c ¼ chigh in one pairing ½ðclow; chighÞ ¼ ð�1; 0Þ�and also as c ¼ clow in another pairing ½ðclow; chighÞ ¼ð0; 1Þ�. A comparison of performance in these two kinds

of trials yields no suggestion of an effect of context. For

the ‘‘different statistics’’ trials (Fig. 9(B), middle panel

and Fig. 9(B), lower panel), there is a clear dependenceof fraction correct on the c-value of the target. Regres-

sion analysis shows that this dependence is predomi-

nantly or exclusively related to the c-value of the target

in S2 (regression slope 0.115, p < 10�4, F -test), not S1(regression slope )0.026, p > 0:1, F -test).

We also analyzed the isodipole experiment with

Dc ¼ 0:4 (Fig. 7) along these lines. Consistent with the

above observations, there was a trend of similar mag-nitude towards greater fraction correct as the c-value ofthe target increased (regression slope 0.040 for the

‘‘same statistics’’ condition, 0.065 for the ‘‘different

statistics’’ condition, each p < 10�2, F -test). Since the c-values of the target in S1 and S2 were similar ðDc ¼ 0:4Þ,we could not separate the influence of the c-value of thetarget in S1 versus S2.

In sum, superimposed on the effect of whether targetstatistics change, there are additional influences of the

position of the target’s statistics along the range inves-

tigated. This dependence is remarkably different for the

luminance and the isodipole experiments. In the lumi-

nance experiments, in ‘‘same statistics’’ trials, there a

context effect: an increase in fraction correct when the

target is closer to random than the typical array. In the

‘‘different statistics’’ trials, fraction correct is higher for

targets that darken than for targets that brighten. In the

isodipole experiments, there is an increase in fraction

correct when the target is closer to the even end of the

range, whether or not there is a change in statistics, and

there is no effect of context.

5. Discussion

5.1. What kinds of image statistics are used in visual

working memory?

In the experiments reported here, we used a modifi-

cation of the Cornelissen and Greenlee (2000) task to

demonstrate that visual working memory can make use

not only of individual pixel values, but also of the sta-

tistical structure of the arrays. We considered three

kinds of statistical structure: changes in luminance,changes in higher-order local correlations, and the

presence or absence of bilateral symmetry. Only the first

two kinds of statistical structure appeared to be useful as

cues in this visual memory task. Surprisingly, removal or

introduction of bilateral symmetry, despite is apparent

salience and importance in visual tasks (Attneave, 1954;

Tyler, 1995; Wenderoth, 1994), did not influence per-

formance.

5.1.1. Improved performance is not due to stimulus set size

This finding cannot be accounted for by differences in

the stimulus set size induced by the statistical structures

we have introduced. The size of a stimulus set, weighted

by the relative frequencies of the stimuli within the set, is

naturally quantified (on a logarithmic scale) by its en-

tropy. (For a general review of entropy and related

concepts, see Cover and Thomas (1991).) The entropy ofan ensemble of random n� n arrays, in which each

check is assigned independently and with equal proba-

bility to one of two states, is Hrandom ¼ n2 bits, since thereis one bit associated with each check’s assignment. For

an n� n array in which luminance statistics are con-

trolled by the parameter c, the ensemble entropy is

HlumðcÞ

¼ � n2

log 2

1þ c2

log1þ c2

� ��þ 1� c

2log

1� c2

� ��;

ð1Þ

since each of the n2 checks are independently assigned to

two states, with probability 1þc2

and 1�c2. (Note that

Hlumð0Þ ¼ Hrandom.) We have seen (Fig. 5) that a change

of c from 0 to 0.25 yields a significant cue in the working

memory task. As seen from Eq. (1), this is only a modest

reduction in entropy: Hlumð0:25Þ � 0:95Hrandom. On theother hand, a change from a random texture to a

completely symmetric n� n square array of checks fails

to yield a cue to the working memory task (Fig. 2).

Page 10: Visual working memory for image statistics

550 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

However, the ensemble entropy is reduced from Hrandom

by a factor of 2 to Hsymm ¼ n2

2bits, since each check on

one half of the array can be assigned with equal likeli-

hood to one of two states. That is, even though the

symmetry manipulation associated with c ¼ 1 induces a

much greater reduction in randomness than the lumi-

nance manipulation associated with c ¼ 0:25, only the

latter cue is available to improve performance in theworking memory task.

5.1.2. An information-theoretic measure of statistical

difference

Next we determine the extent to which the differences

between the ensembles induced by the change in statis-

tics, rather than the size of the ensembles themselves,

might account for the observed performance. A natural

notion of the differences between two statistical ensem-bles is the Kullback–Leibler divergence (Cover & Tho-

mas, 1991). The Kullback–Leibler divergence DKLðPkQÞis an information-theoretic measure of the extent to

which stimuli drawn from an ensemble P have proba-

bilities that are unanticipated if it is expected that they

were drawn from an ensemble Q. The Kullback–Leibler

divergence has an interpretation in terms of the perfor-

mance of an ideal observer on a task that is conceptuallyrelated to the one we used here: it measures how readily

an ideal observer can determine, by observing sample

arrays from an ensemble P , that these samples did not

come from an ensemble Q. This measure is based only

on the extent to which arrays have unequal probabilities

in the two ensembles, and not on the spatial or geo-

metrical aspects of their structure. (However, we note

that this is not an ideal-observer analysis of the task weused. For any memory task, ideal-observer performance

would be perfect. That is, the Kullback–Leibler diver-

gence does not address how well an ideal observer

should perform on our task, but rather, the magnitude

of the statistical cue.)

The Kullback–Leibler divergence DKLðPkQÞ is de-

fined by

DKLðPkQÞ ¼Xi

pi logpiqi

� �; ð2Þ

where pi is the probability of the ith stimulus in theensemble P , qi is the probability of the ith stimulus in the

ensemble Q, and the sum is over all stimuli. The Kull-

back–Leibler divergence is evidently not symmetric in Pand Q, and is infinite if any stimuli occur in P but not

in Q. For both reasons, it is customary to use a sym-

metrized form of DKLðPkQÞ, the ‘‘resistor average’’

(Johnson, Gruner, Baggerly, & Seshagiri, 2001), to

measure the difference between the distributions P andQ. The resistor-average divergence DRAðPkQÞ is twice

the harmonic mean of DKLðPkQÞ and DKLðQkP Þ, and is

defined by

1

DRAðPkQÞ¼ 1

DKLðPkQÞþ 1

DKLðQkPÞ; ð3Þ

with the understanding that if either Kullback–Leibler

divergence (DKLðPkQÞ or DKLðQkP Þ) is infinite, then the

corresponding term in the above equation is set to zero.

The Kullback–Leibler divergences, and hence the

resistor-average divergences, are readily calculated for

the three series of stimuli used, for any pair of levels of

structure. For the luminance series, the calculation is

most straightforward. The assignments of states of the n2

checks are completely independent. For an ensemble of

arrays Plum characterized by a level of structure cP , theprobabilities of the states are 1þcP

2and 1�cP

2. Correspond-

ingly, in an ensemble of arrays Qlum characterized by a

level of structure cQ, the probabilities of the states are1þcQ2

and1�cQ2. Since these assignments are made indepen-

dently at each of the n2 checks, we find from Eq. (2) that

DKLðPlumkQlumÞ

¼ n2

log 2

1þ cP2

log1þ cP1þ cQ

�þ 1� cP

2log

1� cP1� cQ

�:

ð4Þ

For the isodipole series, the states of the 2n� 1 checks in

the first row and first column are assigned randomly and

with equal probability to the two states, while the

assignment of the interior checks requires ðn� 1Þ2independent choices biased by the value of cP or cQ. Thisleads to

DKLðPisodipolekQisodipoleÞ

¼ ðn� 1Þ2

log 2

1þ cP2

log1þ cP1þ cQ

�þ 1� cP

2log

1� cP1� cQ

�:

ð5Þ

Finally, for the symmetry series, the states of the n2

2

checks in one half of the array are assigned at random,

while the assignment of the checks on the opposite half

of the array requires n2

2independent choices biased by

the value of cP or cQ. Thus,

DKLðPsymmkQsymmÞ

¼ n2

2 log 2

1þ cP2

log1þ cP1þ cQ

�þ 1� cP

2log

1� cP1� cQ

�:

ð6Þ

5.1.3. Does performance depend on degree of statistical

difference?

The above Eqs. (4)–(6), combined with the symme-

trization of Eq. (3), yields a natural (but purely statis-

tical) measure of the extent to which each of the pairs of

ensembles differ. In Fig. 10, we compare this measurewith the experimental findings. The divergence between

the fully symmetric and fully random arrays was larger

than that between all of the pairings of the luminance

Page 11: Visual working memory for image statistics

-1 0 1

2

4

6

structure parameter c

sqrt(

K-L

dive

rgen

ce b

etw

een

stim

ulus

pai

rs)

0-0.5 0.5

luminance isodipole A isodipole Bsymmetry

Fig. 10. A summary of the Kullback–Leibler divergences between the

ensemble pairings studied in Experiment II (luminance and isodipole)

and Experiment I (symmetry). For each pair of c-values studied, the

ordinate shows the square root of the resistor-average divergences, Eq.

(3), calculated from Eqs. (4)–(6), with cP ¼ clow and cQ ¼ chigh. Theabscissa is the mean c-value, clowþchigh

2. The asterisks mark the pairings

for which a significant difference in performance between performance

in ‘‘same statistics’’ and ‘‘different statistics’’ trials was observed. For

the luminance series (Fig. 5), Dc ¼ jchigh � clowj ¼ 0:25. For isodipole

series A of Experiment II (Fig. 7), Dc ¼ 0:4. For isodipole series B of

Experiment II (Fig. 8) and the symmetry experiment (Fig. 2(C)),

Dc ¼ 1.

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 551

series with Dc ¼ 0:25, but only the latter pairings pro-

vided a cue in the working memory task. Additionally,

all of the Kullback–Leibler divergences in the isodipole

pairings are substantially larger than those in the lumi-

nance series, but only the pairings in Experiment IIB

ðDc ¼ 1Þ and the most ‘‘even’’ pairing of Experiment

IIA lead to improved performance in the ‘‘different

statistics’’ condition. Finally, although many of theKullback–Leibler divergences in the isodipole pairings

of Experiment IIB are smaller than those of the sym-

metry experiment, only the former reveal an effect of

image statistics.

Much as with ideal observer analyses in other con-

texts (e.g., (Geisler, 1984)), the discrepancies between

our observations and those anticipated from the Kull-

back–Leibler analysis reveal the limitations (and strat-egies) of the visual system. The above analysis shows

that the spatial arrangement of the correlations, and not

merely their statistical strength, is crucial for the repre-

sentation in visual working memory. This is entirely in

parallel with the role of image statistics in texture seg-

mentation and discrimination (Julesz, 1981a, 1981b;

Julesz et al., 1973, 1978; Victor & Conte, 1991). Differ-

ences in power spectra (second-order spatial correlationstructure) are potent cues to texture discrimination and

segmentation. However, only very specific higher-order

spatial correlations (as manifest by isodipole textures)

can support these processes efficiently.

Even within the isodipole texture series, the percep-

tual consequences of a statistical change do not corre-

spond to the Kullback–Leibler divergence. In the

isodipole experiments of Figs. 7 and 8, the ‘‘even’’ end of

the range (c near 1) showed a larger effect of a change in

image statistics than the ‘‘odd’’ end of the range (c near

)1). This asymmetry is not manifest in the Kullback–

Leibler divergences of Fig. 10, which are independent of

the sign of c. The greater salience of visual structure forpositive values of c compared to negative values ob-served here corresponds to the larger size of the visual

evoked potential elicited by interchange between even

and random isodipole textures, compared with that

elicited by interchange between odd and random isodi-

pole textures (Victor & Conte, 1989a). We suspect that

this asymmetry reflects the fact that the arrays at the

even end of the range contain large homogenous patches

(blobs), as well as extended contours (Victor & Conte,1989a, 1989b, 1991)––features that are important in

early visual processing (Field, Hayes, & Hess, 1993;

Julesz, 1991; Kovacs & Julesz, 1993).

On the other hand, the Kullback–Leibler analysis may

account for the interaction of n2, the number of checks in

the array and the size of the cue due to array statistics. As

seen from Fig. 2, overall performance is better for the 16-

element arrays than for the 64-element arrays for allstatistical modalities studied. This main effect of array

size could have multiple explanations, including a re-

duced load on a pixel-based mechanism. More signifi-

cantly, in the twomodalities for which there is a difference

in performance between the ‘‘same statistics’’ and ‘‘dif-

ferent statistics’’ (Fig. 2(A) and (B)), there is less of a

difference for the smaller arrays. This is anticipated from

statistical considerations, since (as seen fromEqs. (4)–(6))for smaller arrays (smaller n), it is more difficult to

determine the ensemble from which an array is drawn. A

ceiling effect may also contribute to the smaller effect of

the statistical cue, but this is unlikely to be the full

explanation, since performance in the isodipole experi-

ments, even at the smaller array size (Fig. 2(B)), wasworse

than that in the luminance experiments (Fig. 2(A)).

5.1.4. Possible alternative explanations

The main message of the above analysis is that it

allows for a comparison across statistical classes on an

equal footing. Not surprisingly, the difference in lumi-

nance statistics needed to provide a cue to visualworking memory is much smaller than the difference in

isodipole statistics. But, given the salience of symmetry

in other contexts (Attneave, 1954; Tyler, 1995; Wende-

roth, 1994), it is particularly striking that a maximal

difference in symmetry (and one which, on a statistical

basis, is substantially larger than the threshold difference

for isodipole statistics) provides no cue whatsoever.

One possibility that might account for this result isthat symmetry is not fully processed within the con-

straints of our stimulus presentation (four simultaneous

targets presented for 600 ms in S1) and interstimulus

Page 12: Visual working memory for image statistics

552 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

interval (200 ms). Recent experiments support a con-

tributing role for this factor: identification of a sym-

metric array is improved when the arrays are presented

in RSVP fashion (Conte, Purpura, & Victor, 2002)

compared to simultaneous presentation, or, when the

interstimulus interval is increased (Conte & Victor,

2003)––though substantial differences between detection

of symmetry and the other kinds of statistics persist evenwith processing times up to 1 sec. Along with the con-

straints implied by symmetry detection on multiple axes

(Wenderoth, 1994) and the distinctive scaling behavior

of symmetry perception (Rainville & Kingdom, 1999,

2002; Tyler, 1999), the slow dynamics of symmetry

analysis compared with the more rapid analysis that

suffices to extract luminance and isodipole statistics

suggests that processing of symmetry is not based on acascading of signals through a hardwired circuit, but

more likely, reflects an iterative hypothesis-testing visual

routine (Hayhoe, 2000).

The distinctive behavior of symmetry cannot be fully

accounted for by detection differences. We recently

performed (Victor, Hardy, & Conte, 2002) detection

experiments based on a similar stimulus set, but a

shorter presentation time (100 ms). Performance levelsattained at jDcj ¼ 1 for symmetry (�40% correct) were

typically achieved for isodipole statistical changes of

jDcj ¼ 0:4, and luminance changes of jDcj < 0:1. Thus,detection differences may account for the much smaller

value of jDcj that is required for an improvement on the

memory task for luminance changes, as compared with

isodipole statistical changes. However, a change in

symmetry ðjDcj ¼ 1Þ that does not lead to an improve-ment on the memory task (Experiment I) is, by this

measure, as detectable as a change in isodipole statistics

ðjDcj ¼ 0:4Þ, that does lead to an improvement in per-

formance (Experiment II). The possibility that detection

differences coupled with differences in processing

dynamics combine to account for what appears to be a

difference in visual working memory is unlikely. This is

because with the longer presentation times used here,detection differences are less prominent.

5.2. Relation to models of texture segregation

Many computational models for texture segregationhave been proposed, with generally similar structure: an

initial local stage, usually consisting of Gabor-like ele-

ments and perhaps local non-linear processing followed

by a second stage, in which local signals are pooled,

perhaps also in a non-linear fashion (Bergen & Adelson,

1988; Chubb & Landy, 1991; Graham, 1989; Graham,

Beck, & Sutter, 1992; Grossberg & Mingolla, 1985;

Malik & Perona, 1990; Victor & Conte, 1991; Wilson,1993; Zhu et al., 1998). Such models have been suc-

cessful in accounting for a wide range of texture dis-

crimination phenomena.

In texture discrimination studies, performance is as-

sayed by asking the subject to segment an image. This is,

fundamentally, a statistical task: the outputs of the local

filters must be pooled in order to identify gradients or

discontinuities. Here, the subject is asked to determine

whether an image has changed. Presumably, the same

early visual mechanisms (‘‘local filters’’) are used in both

tasks. In principle, the outputs of these local filters couldbe retained individually, rather than collectively. Were

this the case, there would be no difference in perfor-

mance between conditions in which the statistical

structure changed, and in which it did not––since there

is the same degree of local change in each case. Thus,

our finding that certain kinds of overall statistical

structure provide a cue in a visual memory task indicates

that the pooled signal is the basis not only for spatialcomparisons, but also for comparisons across time.

5.3. Texture is represented continuously, not categorically

In many sensory and perceptual domains, a stimulusspace that spans a continuum is represented in terms of

discrete categories. Among non-visual domains, experi-

mental evidence for categorical perception has been

found in the processing of phonemes (Aaltonen, Niemi,

Nyrke, & Tuhkanen, 1987; Pisoni & Lazarus, 1974) and

somatosensation (Romo, Merchant, Zainos, & Her-

nandez, 1997). Within vision, evidence for categorical

perception has been seen for color (Amano et al., 2002;Berlin & Kay, 1969; Bornstein & Korda, 1984; Wandell,

1985), orientation (Rosielle & Cooper, 2001), facial

expression (Roberson & Davidoff, 2000), and animal

form (Freedman, Riesenhuber, Poggio, & Miller, 2001,

2002).

Consider a discrimination task within a domain

parameterized by a parameter c, in which the subject is

to discriminate stimuli A and B (characterized, respec-tively, by cA or cB). In this setting, the hallmark of a

categorical representation (Aaltonen et al., 1987; Born-

stein & Korda, 1984; Friedman-Hill, Robertson, &

Treisman, 1995; Pisoni & Lazarus, 1974; Wandell, 1985)

is that there is a jump in fraction correct and a decrease

in reaction time when cA and cB straddle a category

boundary, compared to performance with the same

Dc ¼ jcA � cBj when cA and cB are within the same cat-egory. Conversely, with a graded representation, per-

formance depends primarily on jcA � cBj, but not on the

particular values of cA or cB.One might speculate that as indices of texture-like

surface properties, image statistics may play a role

similar to that of color (Cho et al., 2000; Harvey &

Gervais, 1981). However, our results provide no support

for a categorical representation of texture, either by thefraction correct or reaction time criteria. While the data

of Fig. 7 ðDc ¼ 0:4Þ raise the possibility that the pure

even ðc ¼ 1Þ texture is in a category by itself, and that all

Page 13: Visual working memory for image statistics

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 553

statistics corresponding to c < 1 are equivalent, the data

of Fig. 8 ðDc ¼ 1Þ show that this is a manifestation of a

threshold. One possible basis of this difference is that

categorical perception derives not from a functional role

in surface (Cho et al., 2000; Harvey & Gervais, 1981) or

object (Freedman et al., 2001, 2002; Rosielle & Cooper,

2001) identification, but rather, arises from a graded

stimulus domain via verbal coding or storage of stimuli(Roberson & Davidoff, 2000). The latter authors

showed, in a visual memory task, that verbal interfer-

ence removed the hallmarks of categorical perception,

both for color and facial expression. The rather abstract

nature of our stimuli may have precluded such verbal

processes, and thus, revealed an underlying graded

representation.

Acknowledgements

Portions of this work were presented at the 2001

meeting of the Society for Neuroscience and the 2002

meeting of the Association for Research in Vision and

Ophthalmology, Ft. Lauderdale, FL, and was supported

by NIH NEI EY7977. We thank Caitlin Hardy for

assistance with data collection and Jeff Tsai for pro-gramming assistance.

Appendix A. Details of stimulus construction

Here we describe the construction of stimulus arrays

across the range of the structure parameter c, for the

three kinds of statistical structure studied: luminance,isodipole textures, and symmetry. We have two goals:

first, generation of an array that has a prescribed value

of c (for use in S1), and second, generation of a second

array (for use in S2) that has a second prescribed value

of c, and in which only a prescribed number of checks

have changed in luminance.

A.1. Luminance statistics

For the luminance experiments, the value of c deter-

mines the number of checks that are white ð1þc2Þ and the

number of checks that are black ð1�c2Þ. Since the total

number of checks must be an integer ðNÞ, these ratios

can only achieve certain discrete values. For the exper-

iments described here, we only used values of c for

which these ratios could be achieved exactly––that is,

values of c from )1 to 1, in steps of 2N. Once c is specified,

the number of white and black checks is then specified,

as Nw ¼ Nð1þc2Þ and Nb ¼ Nð1�c

2Þ, respectively. Arrays

were then constructed by random placement of therequisite number of white and black checks.

If k of the N checks change in luminance between S1

and S2, the maximum increase in c that can occur is 2kN .

This occurs if all k of the altered checks change from

black to white. Similarly, the maximal decrease in c thatcan occur is a change of � 2k

N , which happens if all kaltered checks change from white to black. More gen-

erally, changing the state of k2þ NDc

4checks from black to

white and k2� NDc

4checks from white to black (a total of k

checks) leads to a net change in c of Dc.For all of the experiments described here, the above

quantities are non-negative integers. To create the S2

array from the S1 array, the number of black checks and

white checks to be flipped in contrast is determined by

the above rules, and their location is determined at

random from the location of the Nb black checks and the

Nw white checks in S1.

A.2. Isodipole statistics

For the other kinds of statistics, the goal was the

same, but the details differ. For the arrays based on the

even and odd isodipole textures (Fig. 2(B)), we pro-ceeded as follows. In an even ðc ¼ 1Þ or odd ðc ¼ �1Þisodipole texture array (Julesz et al., 1978), the state ai;j(+1 or )1) of the check in the ith row and jth column is

forced to obey the recursion rule

ai;jai�1;jai;j�1ai�1;j�1 ¼ c: ðA:1ÞThis rule, along with a random assignment of the states

of checks in the initial row ða1;jÞ and initial column ðai;1Þ,generates a texture in which half of the checks (on

average) are in either state, and in which there are no

pairwise or third-order correlations. Isodipole textures

with intermediate values of c were constructed accordingto the ‘‘propagated decorrelation’’ textures of Victor

(1985). For these textures, the deterministic rule (A.1) is

replaced by the probabilistic rule

probfai;jai�1;jai;j�1ai�1;j�1 ¼ 1g ¼ cþ 1

2: ðA:2Þ

Note that since the value of the quadruple product in the

above expression must be either +1 or )1, Eq. (A.2)

reduces to Eq. (A.1) for c ¼ 1 or c ¼ �1. For interme-

diate values of c, the average value of the product in Eq.

(A.2) (over an infinite sample of the texture) is c.For an n� n square array of N elements, there are

ðn� 1Þ2 instances at which the probabilistic rule (A.2)

can be applied: namely, all ði; jÞ pairs for which 26 i6 nand 26 j6 n. Consequently, we replace the probabilisticrule of Eq. (A.2) by the deterministic rule

1

ðn� 1Þ2Xn

i¼2

Xn

j¼2

ai;jai�1;jai;j�1ai�1;j�1 ¼ c: ðA:3Þ

Again, since the value of the quadruple product in theabove expression must be either +1 or )1, this averagecan only have specific values, ranging from )1 to +1, in

steps of 2

ðn�1Þ2. To construct an array corresponding to a

Page 14: Visual working memory for image statistics

554 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

particular value of c that is intermediate between these

achievable values, we used either the next-highest ex-

actly achievable value, cabove, or the next-lowest exactly

achievable value, cbelow, to determine exactly how many

of the recursion products in Eq. (A.3) are equal to +1,

and exactly how many are equal to )1. We select either

cabove or cbelow with probabilities pabove and pbelow to en-

sure that the average value of the recursion products inEq. (A.3) is exactly c. That is,

pabove ¼c� cbelow

cabove � cbelowand

pbelow ¼ cabove � ccabove � cbelow

: ðA:4Þ

For the isodipole textures, changing the state of kchecks can alter the value of c in Eq. (A.3) by as much as

8kðn�1Þ2, since each of the k checks can participate in up to

four products, each of which can have the value 1

ðn�1Þ2 or

� 1

ðn�1Þ2. However, since the checks involved in these

quadruple products overlap, it is not straightforward to

construct a pair of arrays that have prescribed values of

c, and which differ at exactly k checks. The first step in

doing this was to construct two arrays with the desired

values of c without any constraints on the number of

checks at which they differed. Since these arrays were

constructed independently, they typically mismatched at

approximately N2of the checks. Thus, the typical number

of mismatches in the independently constructed arrays is

larger than the number of desired mismatches, k. Then,we sought to apply a sequence of flips of entire rows

and/or columns of checks to one of the arrays, so that

the number of mismatched checks would equal k. Note

that flipping entire rows or columns does not change the

value of c, since these transformations invert the state of

two checks in every affected quadruple product. Thissearch combined a Monte Carlo strategy and strategies

used in the solution of Berlekamp’s switching game

(Fishburn & Sloane, 1989): essentially, a game whose

goal is to minimize the number of mismatches. We then

sought (by iterating this strategy) to combine these pairs

into quadruples of arrays ðA1;A2;B1;B2Þ such that (i) A1

and A2 corresponded to cA, (ii) B1 and B2 corresponded

to cB, (iii) the pairs ðA1;A2Þ, ðA2;B1Þ, and ðB1;B2Þ eachdiffered by k flips. A library of such quadruples was

created via repeating this search procedure. The library

was further enlarged by (a) flipping all the arrays within

a quadruple along their horizontal and/or vertical axes,

and (b) randomly multiplying all the arrays within a

quadruple by a randomly chosen even ðc ¼ 1Þ array.

These transformations preserve the conditions (i), (ii),

and (iii) above, and also ensured that the positions of thechecks that were flipped between S1 and S2 were sym-

metrically distributed, and that the stimuli themselves

were approximately luminance-balanced.

To generate the stimuli for individual trials, each ar-

ray in S1 was drawn from the middle elements (A2 and

B1) of a quadruple randomly selected from this library.

(Each array was drawn from an independently con-

structed quadruple.) To create the S2 target in a ‘‘dif-

ferent statistics’’ trial, an A2 array in S1 is replaced in S2

by the array B1 drawn from the same quadruple, or a B1

array in S1 is replaced in S2 by the A2 array from the same

quadruple. To create the S2 target in a ‘‘same statistics’’

trial, an A2 array in S1 is replaced by an A1 array in S2, or

a B1 array in S1 is replaced by a B2 array in S2.

The reason for the elaborate construction described

above is that it ensures that the exemplar chosen for an

array in S1 gives no indication as to whether it is a

target, nor whether the trial will be a ‘‘different statis-

tics’’ trial or not. It leaves open the possibility that someoverall difference between the A1 arrays and the A2 ar-

rays, or between the B1 arrays and the B2 arrays, might

provide an extra cue in S2 for the ‘‘same statistics’’ trial

that does not require comparison with S1. However, this

kind of artifact would produce the opposite of the result

we observed. Additionally, such cues could not be

identified by the investigators, who had full knowledge

of the construction, even after extended scrutiny of thetextures. We also verified that the changed and un-

changed checks were approximately uniformly distrib-

uted throughout the arrays, and that nearly all arrays

were within 1 or 2 checks of being perfectly luminance-

balanced, even though we did not explicitly balance

luminance.

A.3. Symmetry

We constructed the arrays based on symmetry (Fig.

2(C)–(E)) as follows. An array with perfect ðc ¼ 1Þtwofold symmetry (with either a horizontal or a verticalmirror axis) can be constructed by randomly assigning

the luminance values to the N2checks in one half of the

array. Each of these checks is then paired (via symme-

try) with another check, and the luminance assigned to

this paired check must match to preserve the symmetry.

Graded amounts of symmetry correspond to an array in

which a fraction ðcþ12Þ of the paired checks match, and

the rest of the pairs mismatch. Thus, for c ¼ 0, exactlyhalf of the pairs match, for 0 < c < 1, more pairs match

than would be expected by chance, and for c ¼ �1, all of

the paired checks mismatch. Changing the state of kchecks can change the match versus mismatch state of

up to k of the N2pairs, and thus, change the value of c by

amounts ranging from � 4kN to 4k

N (in steps of 8N) in either

direction. Thus, to change array statistics from perfect

mirror symmetry ðc ¼ 1Þ to mirror symmetry withluminance inversion ðc ¼ �1Þ, as in Fig. 2(E) (corre-

sponding to Dc ¼ 2), half of the checks in the target

array must be flipped between S1 and S2. This cannot be

Page 15: Visual working memory for image statistics

J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556 555

accomplished with N ¼ 64 checks and k ¼ 16 flips, but

only with the smaller array size (N ¼ 16 checks, k ¼ 8

flips).

To create arrays in which changing k checks resulted

in a switch from vertical symmetry to horizontal sym-

metry (Fig. 2(D)), we randomly assigned luminance

values to the upper left quadrant of the array (N4checks).

Each of these checks was used to determine the value ofthe three other checks that were related by horizontal

and vertical mirror operations. Quadruples of checks

with configurations � þ1 þ1

þ1 þ1

� �contributed 4

N to both

vertical and horizontal symmetry; quadruples with

configurations � þ1 �1�1 þ1

� �contributed � 4

N to both

vertical and horizontal symmetry; quadruples with

configurations � þ1 þ1

�1 �1

� �contributed 4

N to vertical

symmetry but � 4N to horizontal symmetry; quadruples

with configurations � þ1 �1

þ1 �1

� �contributed � 4

N to

vertical symmetry but 4N to horizontal symmetry; and the

eight other quadruples with configurations like

� þ1 �1

þ1 þ1

� �and its rotations contributed 0 to vertical

and horizontal symmetry. By choosing appropriate

fractions of these configurations in S1 and S2, the de-

sired amounts of symmetry and number of checks that

were flipped were obtained.

All of the trials made use of values of c (0, 1 and )1)that could be achieved exactly. Moreover, since N was a

multiple of 16 and k was a multiple of 8, it was possible

to arrange the luminance assignments so that S1 and S2

were both precisely luminance-balanced, for the stimuli

involving only vertical symmetry (Fig. 2(C) and (E)) as

well as those with the additional constraint of horizontal

symmetry (Fig. 2(D)).

References

Aaltonen, O., Niemi, P., Nyrke, T., & Tuhkanen, M. (1987). Event-

related brain potentials and the perception of a phonetic contin-

uum. Biological Psychology, 24(3), 197–207.

Amano, K., Uchikawa, K., & Kuriki, I. (2002). Characteristics of color

memory for natural scenes. Journal of the Optical Society of

America A––Optics Image Science and Vision, 19(8), 1501–1514.

Attneave, F. (1954). Some informational aspects of visual perception.

Psychological Review, 61, 183–193.

Bergen, J. R., & Adelson, E. H. (1988). Early vision and texture

perception. Nature (333), 363–364.

Berlin, B., & Kay, P. (1969). Basic color terms: Their universality and

evolution. Berkeley: University of California Press.

Bornstein, M. H., & Korda, N. O. (1984). Discrimination and

matching within and between hues measured by reaction times:

Some implications for categorical perception and levels of infor-

mation processing. Psychological Research, 46(3), 207–222.

Caelli, T., & Julesz, B. (1978). On perceptual analyzers underlying

visual texture discrimination: Part I. Biological Cybernetics, 28(3),

167–175.

Caelli, T., Julesz, B., & Gilbert, E. (1978). On perceptual analyzers

underlying visual texture discrimination: Part II. Biological Cyber-

netics, 29(4), 201–214.

Cho, R. Y., Yang, V., & Hallett, P. E. (2000). Reliability and

dimensionality of judgments of visually textured materials. Per-

ception & Psychophysics, 62(4), 735–752.

Chubb, C., & Landy, M. (1991). Orthogonal distribution analysis: A

new approach to the study of texture perception. In M. S. Landy &

J. A. Movshon (Eds.), Computational models of visual processing

(pp. 291–301). Cambridge, MA: MIT Press.

Conte, M. M., Purpura, K. P., & Victor, J. D. (2002). Processing of

image symmetry in an RSVP task. Society for Neuroscience, 28,

Orlando, FL.

Conte, M. M., & Victor, J. D. (2004). Temporal stability of image

statistics in visual working memory. Vision Sciences Society 2003,

Sarasota, FL.

Cornelissen, F. W., & Greenlee, M. W. (2000). Visual memory for

random block patterns defined by luminance and color contrast.

Vision Research, 40(3), 287–299.

Cover, T. M., & Thomas, J. A. (1991). Elements of information theory.

In Wiley series in telecommunications (p. 542). New York: Wiley.

Field, D. J., Hayes, A., & Hess, R. F. (1993). Contour integration by

the human visual system: Evidence for a local ‘‘association field’’.

Vision Research, 33(2), 173–193.

Fishburn, P. C., & Sloane, N. J. A. (1989). The solution to

Berlekamp’s switching game. Discrete Mathematics, 74, 263–290.

Freedman, D. J., Riesenhuber, M., Poggio, T., & Miller, E. K. (2001).

Categorical representation of visual stimuli in the primate

prefrontal cortex. Science, 291(5502), 312–316.

Freedman, D. J., Riesenhuber, M., Poggio, T., & Miller, E. K. (2002).

Visual categorization and the primate prefrontal cortex: Neuro-

physiology and behavior. Journal of Neurophysiology, 88(2), 929–

941.

Friedman-Hill, S. R., Robertson, L. C., & Treisman, A. (1995).

Parietal contributions to visual feature binding: Evidence from a

patient with bilateral lesions. Science, 269(5225), 853–855.

Geisler, W. S. (1984). Physical limits of acuity and hyperacuity. Journal

of the Optical Society of America A, 1(7), 775–782.

Graham, N. (1989). Visual pattern analyzers. Oxford: Clarendon

Press.

Graham, N., Beck, J., & Sutter, A. (1992). Nonlinear processes in

spatial-frequency channel models of perceived texture segregation:

Effects of sign and amount of contrast. Vision Research, 32(4), 719–

743.

Grossberg, S., & Mingolla, E. (1985). Neural dynamics of perceptual

grouping: Textures, boundaries, and emergent segmentations.

Perception & Psychophysics, 38(2), 141–171.

Harvey, L. O., Jr., & Gervais, M. J. (1981). Internal representation of

visual texture as the basis for the judgment of similarity. Journal of

Experimental Psychology––Human Perception and Performance,

7(4), 741–753.

Hayhoe, M. (2000). Vision using routines: A functional account of

vision. Visual Cognition, 7(1/2/3), 43–64.

Johnson, D. H., Gruner, C. M., Baggerly, K., & Seshagiri, C. (2001).

Information-theoretic analysis of neural coding. Journal of Com-

putational Neuroscience, 10(1), 47–69.

Joseph, J. S., & Optican, L. M. (1996). Involuntary attentional shifts

due to orientation differences. Perception & Psychophysics, 58(5),

651–665.

Julesz, B. (1981a). Textons, the elements of texture perception, and

their interactions. Nature, 290(5802), 91–97.

Julesz, B. (1981b). A theory of preattentive texture discrimination

based on first-order statistics of texture. Biological Cybernetics (41),

131–138.

Page 16: Visual working memory for image statistics

556 J.D. Victor, M.M. Conte / Vision Research 44 (2004) 541–556

Julesz, B. (1991). Early vision and focal attention. Reviews of Modern

Physics, 63(3), 735–772.

Julesz, B., Gilbert, E. N., Shepp, L. A., & Frisch, H. L. (1973).

Inability of humans to discriminate between visual textures that

agree in second-order statistics––revisited. Perception, 2(4), 391–

405.

Julesz, B., Gilbert, E. N., & Victor, J. D. (1978). Visual discrimination

of textures with identical third-order statistics. Biological Cyber-

netics, 31(3), 137–140.

Kovacs, I., & Julesz, B. (1993). A closed curve is much more than an

incomplete one: Effect of closure in figure-ground segmentation.

Proceedings of the National Academy of Sciences USA, 90(16),

7495–7497.

Malik, J., & Perona, P. (1990). Preattentive texture discrimination with

early vision mechanisms. Journal of the Optical Society of America

A, 7(5), 923–932.

Marr, D. (1982). Vision. New York: W.H. Freeman & Co.

Pisoni, D. B., & Lazarus, J. H. (1974). Categorical and noncategorical

modes of speech perception along the voicing continuum. Journal

of the Acoustical Society of America, 55(2), 328–333.

Purpura, K. P., Victor, J. D., & Katz, E. (1994). Striate cortex

extracts higher-order spatial correlations from visual textures.

Proceedings of the National Academy of Sciences USA, 91(18),

8482–8486.

Rainville, S. J., & Kingdom, F. A. (1999). Spatial-scale contribution to

the detection of mirror symmetry in fractal noise. Journal of the

Optical Society of America A––Optics Image Science and Vision,

16(9), 2112–2123.

Rainville, S. J., & Kingdom, F. A. (2002). Scale invariance is driven by

stimulus density. Vision Research, 42(3), 351–367.

Roberson, D., & Davidoff, J. (2000). The categorical perception of

colors and facial expressions: The effect of verbal interference.

Memory & Cognition, 28(6), 977–986.

Romo, R., Merchant, H., Zainos, A., & Hernandez, A. (1997).

Categorical perception of somesthetic stimuli: Psychophysical

measurements correlated with neuronal events in primate medial

premotor cortex. Cerebral Cortex, 7(4), 317–326.

Rosielle, L. J., & Cooper, E. E. (2001). Categorical perception of

relative orientation in visual object recognition. Memory &

Cognition, 29(1), 68–82.

Treisman, A. (1982). Perceptual grouping and attention in visual

search for features and for objects. Journal of Experimental

Psychology––Human Perception and Performance, 8(2), 194–214.

Tyler, C. W. (1995). Empirical aspects of symmetry perception. Spatial

Vision, 9(1), 1–7.

Tyler, C. W. (1999). Human symmetry detection exhibits reverse

eccentricity scaling. Visual Neuroscience, 16(5), 919–922.

Victor, J. D. (1985). Complex visual textures as a tool for studying the

VEP. Vision Research, 25(12), 1811–1827.

Victor, J. D. (1986). Isolation of components due to intracortical

processing in the visual evoked potential. Proceedings of the

National Academy of Sciences USA, 83(20), 7984–7988.

Victor, J. D., & Conte, M. M. (1989a). Cortical interactions in texture

processing: Scale and dynamics. Visual Neuroscience, 2(3), 297–

313.

Victor, J. D., & Conte, M. M. (1989b). What kinds of high-order

correlation structure are readily visible? Investigative Ophthalmol-

ogy & Visual Science, 30(Suppl.), 254.

Victor, J. D., & Conte, M. M. (1991). Spatial organization of nonlinear

interactions in form perception. Vision Research, 31(9), 1457–1488.

Victor, J. D., & Conte, M. M. (1996). The role of high-order phase

correlations in texture processing. Vision Research, 36(11), 1615–

1631.

Victor, J. D., & Conte, M. M. (2001). Dynamics of selective spatial

attention in a working memory task. Association for Research in

Vision and Ophthalmology, 42 (p. 863). Ft. Lauderdale, FL.

Victor, J. D., Hardy, C., & Conte, M. M. (2002). Visual processing of

image statistics: Qualitative differences between local and global

statistics; quantitative differences between low- and high-order

statistics. Vision Sciences Society 2002, Sarasota, FL.

Wandell, B. A. (1985). Color measurement and discrimination. Journal

of the Optical Society of America A, 2(1), 62–71.

Wenderoth, P. (1994). The salience of vertical symmetry. Perception,

23(2), 221–236.

Wilson, H. R. (1993). Nonlinear processes in visual pattern discrim-

ination. Proceedings of the National Academy of Sciences USA,

90(21), 9785–9790.

Zhu, S. C., Wu, Y., & Mumford, D. (1998). Filters, random fields and

maximum entropy (FRAME): Towards a unified theory for texture

modeling. International Journal of Computer Vision, 27(2), 107–126.